Antonio
García Alcaraz
a,
Alicia
Rey Planells
ab,
Arturo
Espinosa Ferao
*a and
Rainer
Streubel
*c
aDepartamento de Química Orgánica, Facultad de Química, Universidad de Murcia, Campus de Espinardo, 30071 Murcia, Spain. E-mail: artuesp@um.es
bFaculty of Pharmacy. University of Castilla-La Mancha, Calle Almansa 14 – Edif. Bioincubadora, 02008 Albacete, Spain
cInstitut für Anorganische Chemie. Rheinische Friedrich-Wilhelms-Universität Bonn, Gerhard-Domagk-Str. 1, 53121 Bonn, Germany. E-mail: r.streubel@uni-bonn.de
First published on 7th January 2025
Compared to aziridines, azaphosphiridines, which formally result from the replacement of a carbon atom by phosphorus, have been much less studied. In this work, accurate values for one of the most prominent properties, the ring strain energy (RSE), have been theoretically examined for a wide range of azaphosphiridine derivatives. Strongly related aspects of interest for developing the use of azaphosphiridines in heteroatom and polymer chemistry are ring opening reactions and polymerisations, the latter facilitated by their significantly high RSE. While methyl groups have little influence on the RSE, complexation with different metal moieties increases the RSE in all cases, and an increase was also found upon oxidation to the corresponding P-oxides and other σ5λ5-P derivatives. The highest RSE was found for the P-protonated azaphosphiridinium cation and azaphosphiridines with exocyclic double bonds. A correlation of the RSEs with the relaxed force constants of the endocyclic ring bonds and AIM-derived parameters in the ring critical points, such as the electron density, ρ(r), and the Lagrangian of the kinetic energy, G(r), was found. A relatively low barrier to P–C bond cleavage via nucleophilic attack of MeNH2 on phosphorus points to the possibility of ring-opening polymerisation.
![]() | ||
Fig. 1 Three-membered heterocycles I–VII (E = O, S, NR; [M] = BR3, M(CO)5; M = Cr, Mo, W; external lines denote ubiquitous substituents). |
On the other hand, the incorporation of a NH unit in phosphirane II results in the formation of σ3λ3-azaphosphiridines III.6,7 In this context, σ refers to the coordination number, while λ refers to the number of electrons involved in bonding at the phosphorus atom. These compounds are fascinating in the realm of phosphorus-based heterocycles due to their potential reactivity based on the three distinct degrees of polarisation in their endocyclic bonds and significantly increased ring strain energy (RSE) for III8 (vide infra). These characteristics render III intriguing candidates for potential applications in ring-opening polymerisations.
The chemistry of uncomplexed or higher P-coordination index azaphosphiridine made significant advances in the 1970s and 1980s.9 In particular, in 1973 the first σ5λ5-azaphosphiridines IV were successfully synthesised by reacting the azine of hexafluoroacetone with different phosphorus derivatives.10 Subsequently, Niecke reported the synthesis of the σ4λ5-azaphosphiridine P-imine V using the reaction between diazomethane and a bis(imino)phosphorane.11 The synthesis of σ4λ5-azaphosphiridine P-sulphides V was also achieved using two different protocols.12 In the first, a thioxophosphorane was employed, which underwent a reaction with tert-butyl azide, resulting in the formation of the azaphosphiridine through photochemically induced nitrogen extrusion. In the second approach, the synthesis of V is achieved through the reaction of a phosphorane and diazomethane, with a thermally promoted nitrogen removal step. The first free σ3λ3-azaphosphiridine III was also prepared by Niecke through a thermally induced valence isomerization of an imino(methylene)phosphorane.7 Years later, Majoral devised another strategy for the synthesis of III using two equivalents of lithium amide, which reacted with phosphaalkenes to yield the detectable phosphaguanidines that spontaneously converted to the corresponding heterocycles,13 but no further research was published afterwards. Subsequently, azaphosphiridine P-oxides were proposed as intermediates in the base-induced rearrangement of α-halophosphonamidates. However, in no case could the three-membered heterocycle be isolated.14–17 The development of the chemistry of terminal phosphinidene metal complexes also led to the preparation of azaphosphiridine complexes VI18,19via reaction with imines.20 The development of the Li/Cl phosphinidenoid complex methodology21,22 enabled broader access to new complexed azaphosphiridine complexes through the reaction with imines18 or carbodiimides.23,24 In the latter case, the iminoazaphosphiridine stands out for its remarkable reactivity towards the activation of small molecules, such as phenyl isocyanate, carbon dioxide, carbon monoxide, elemental sulfur or isocyanides.24,25 In particular, the reaction with CO2 was subjected to detailed study, which revealed that it behaves as a masked FLP system and responds to the presence of the reagent as a stimulus.
The only other related P-heterocycle with three different polar bonds that has been the subject of greater investigation are σ3λ3-oxaphosphiranes VII, which also display a high RSE.26 Although compounds III and VII possess a comparable core structure, one of the principal anticipated benefits of azaphosphiridines is their elevated basicity, which is attributed to the presence of nitrogen. This should enhance their nucleophilic character towards both hard and soft electrophilic centres at the N and P ring atoms, respectively. Furthermore, azaphosphiridines are anticipated to exhibit enhanced versatility with regard to functionalisation, thereby facilitating their potential utilisation in subsequent applications.
A computational study of these heterocycles III–VI is essential to further investigate their structural and electronic properties. Lammertsma was first to undertake a theoretical investigation of azaphosphiridines. Utilising isodesmic reactions, he estimated the RSE of the unsubstituted azaphosphiridine to be approximately 26.5 kcal mol−1, which was considerably higher than that of phosphirane (21.4 kcal mol−1) but slightly lower than the RSE of aziridine (28.2 kcal mol−1).27 The aforementioned values were recalculated years later using the more reliable homodesmotic reactions, resulting in 23.7, 19.9 and 27.6 kcal mol−1 for these three rings, respectively.19 This high RSE value prompted the computational study of the ring-opening reactions in the parent heterocycle.27 The cleavage of the endocyclic P–C and C–N bonds are endothermic processes, with the former being slightly more favoured both kinetically and thermodynamically. Conversely, the cleavage of the P–N bond is a slightly exergonic process with a high TS (transition state) which additionally involves a migration of a hydrogen from C to P. Therefore, the only feasible process with an accessible energy barrier is the breaking of the endocyclic P–C bond. Complexation to one unit of W(CO)5 was shown to have negligible effect on the energies of these opening reactions.
Years later, Espinosa Ferao and Streubel conducted a more detailed analysis of the potential energy surfaces of a small but representative set of decomplexed azaphosphiridines,8 and a notable variation in the RSE, increasing from 22.6 to 38.0 kcal mol−1, was observed when moving from the parent compound of III to the oxidized form VI (E = O). One year later, the same authors found that complexation of the azaphosphiridine to one Cr(CO)5 unit did not result in a significant change in the RSE value.8 Furthermore, RSE values of 50.58 and 52.2 kcal mol−1 were found for the E and Z isomers of 3-imino-azaphosphiridine P–W(CO)5, respectively, indicative of a super strained ring that may explain their enhanced reactivity.
Notwithstanding the aforementioned investigations, a significant knowledge gap remained with regard to the effects of substituents. Consequently, a comprehensive study of the RSEs for a diverse range of azaphosphiridines was undertaken to gain further insights. This study systematically examines the RSE of a series of azaphosphiridines with different substituents (Fig. 2), as they are a priori excellent candidates for polymerisation reactions and exhibit analogy with aziridines and with the recently studied oxaphosphiranes.28 Thus, the unsubstituted azaphosphiridine 1a has been included together with those differently substituted with methyl groups in the three positions as well as cationic azaphosphiridinium species. Additionally, complexed azaphosphiridines in which the phosphorus LP (lone pair) is involved in coordinative bonding and, hence, typically protected by Fe(CO)4, Cr(CO)5, Mo(CO)5 and W(CO)5, as well as with borane, have been investigated. Furthermore, oxidised species such as the P-oxide and the pentacoordinated P(V) derivative, as well as functionalisation with different exocyclic unsaturations at carbon, have been included in the study. Finally, the objective is to examine the potential correlation between RSE and the electronic or structural intrinsic properties of the system in order to arrive to some reactivity predictions.
Table 1 presents the calculated RSE values for all compounds. For purposes of comparison, the values obtained at the DLPNO-CCSD(T)/def2-QZVPP(ecp) level and at the B97-3c optimisation level (in brackets) have been included. While the latter are an acceptable estimate, they include in some cases (Fe-1a) absolute variations of up to 2.6 kcal mol−1. Firstly, it can be noted that the level used here (DLPNO-CCSD(T)/def2-QZVPP//B97-3c) is in good agreement with the DLPNO-CCSD(T)/def2-TZVPP//BP86/def2-TZVP level that was used to calculate the RSE reported for the parent azaphosphiridine 1a (23.7 kcal mol−1).8 Furthermore, there is a satisfactory correlation with the estimated value of 26.2 kcal mol−1 derived through the additive method, which considers bond contributions to ring strain (6.18, 11.42 and 8.65 kcal mol−1 for P–C, P–N and C–N bonds, respectively) while the additive estimation using atomic contributions yields a value of 21.9 kcal mol−1 (8.65, 10.22 and 3.07 kcal mol−1 for C, N and P atoms, respectively).32
Compound | RSE | Compound | RSE |
---|---|---|---|
1a | 24.4 (24.8) | Fe-1a | 32.1 (34.7) |
1b | 23.2 (22.9) | Cr-1a | 30.4 (31.3) |
1c | 24.4 (24.1) | Mo-1a | 30.8 (32.0) |
1d | 24.4 (24.0) | W-1a | 31.7 (31.5) |
1e | 53.0 (53.1) | B-1a | 34.3 (34.0) |
1f | 27.8 (27.9) | O-1a | 38.9 (38.6) |
1g | 34.1 (34.3) |
W-1a![]() |
41.04 (40.5) |
1a![]() |
33.8 (32.4) |
W-1a![]() |
45.0 (43.6) |
1a![]() |
37.3 (34.8) |
W-1a![]() |
46.9 (45.4) |
1a![]() |
40.3 (38.1) | 1a·H+ | 23.1 (24.3) |
Secondly, the introduction of a methyl group at either carbon or phosphorus results in a negligible change in the RSE value. However, at nitrogen, a slight decrease in RSE is observed (from 24.4 to 23.2 kcal mol−1). P-complexation with an Fe(CO)4 group has been observed to increase the RSE by almost 8 kcal mol−1, a result that is consistent with that observed for oxaphosphiranes.28 Nevertheless, complexation with Cr(CO)5, Mo(CO)5 and W(CO)5 yields RSE values of 30.4, 30.8 and 31.7 kcal mol−1, respectively, with a maximum value of 34.3 kcal mol−1 observed in the case of complexation with BH3. The oxidation of the parent compound to the P-oxide (σ4λ5-P) O-1a also results in an increase to 38.9 kcal mol−1, which exhibits the same trend reported for oxaphosphiranes.28 In contrast, the azaphosphiridine with a σ5λ5-P centre 1f exhibited a comparatively lower value of 27.8 kcal mol−1. The introduction of exocyclic unsaturations on the ring carbon of the azaphosphiridines, namely CH2,
NH and
O, resulted in a considerable increase in RSE in all cases, reaching values of 41.0, 45.0 and 46.9 kcal mol−1, respectively (Table 1). These results demonstrate a clear trend of
CH2 <
NH <
O, as already reported for oxaphosphiranes.23
P-protonated azaphosphiridinium 1e is the derivative with the highest RSE of all species studied herein, as was the case for the oxaphosphiranium cation for the oxygen analogues.28 This could be due to the absence of the phosphorus lone pair, which would provide the well-known ring strain relaxation mechanism by increasing the ‘p’ character of the atomic orbitals involved in endocyclic bonds.33 However, it is expected that the most basic centre of azaphosphiridine 1a is the N atom, due to the high reluctance of the P atom to acquire the required sp3 hybridisation in the corresponding protonated species. Indeed, at the working level of theory, the P-to-N proton shift from 1e to the N-protonated species 1a·H+ was calculated to be very exergonic (ΔG = −22.8 kcal mol−1), with a moderate barrier (ΔG‡ = 30.6 kcal mol−1) for the intramolecular process, thus suggesting a preferred intermolecular pathway (not computed). The N atom of 1a is expected to be somewhat more basic than the most strained aziridine (reported pKb = 6.02).34 The resulting 1a·H+ derivative displays slightly lower RSE than the corresponding conjugate base 1a (Table 1).
In contrast, the high RSE of the σ5λ5-spiro-azaphosphiridine 1g does not correlate with the particularly low value reported for the related σ5λ5-spiro-oxaphosphirane (9.03 kcal mol−1).35 This can be tentatively explained by the lower apicophilicity of the NH group in comparison to O, which resulted in a slightly distorted square pyramidal geometry for 1g (τ5 = 0.178), in contrast to the trigonal bipyramidal structure observed for the σ5λ5-spiro-oxaphosphirane.
![]() | ||
Fig. 3 Plot of the RSEs (kcal mol−1) versus the relaxed force constants of the C–N bonds. In red are compounds with exocyclic unsaturations and in grey 1e excluded from the correlation. |
It can be observed that compounds with exocyclic double bonds exhibit the highest relaxed force constant for the endocyclic C–N bond. This is attributed to the contribution of resonance structures, which endow them with a partial double bond character. Such a phenomenon is particularly evident in compounds W-1aC, W-1a
NH and W-1a
O (Scheme 2).
![]() | ||
Scheme 2 Resonance structures of azaphosphiridines with exocyclic unsaturations. X = CH2, NH, O; E = LP, W(CO)5. |
Similarly, an acceptable correlation was identified for the endocyclic N–P bonds (R2 = 0.5673) when excluding derivatives 1e and 1g and those with exocyclic unsaturations (Fig. 4). The distinct behaviour of the latter is a logical consequence of the significant delocalisation of the lone pair (LP) of the ring N atom across the exocyclic position in these enamine, amide or amidine groups. This results in the acquisition of a significant electronic deficiency on the endocyclic N atom adjacent to the P atom, due to the participation of the aforementioned resonance structures (Scheme 2). The relative importance of this phenomenon increases with the electronegativity of the exocyclic heavy atom, following the order C < N < O. Similarly, for the endocyclic P–C bonds, an even stronger correlation (R2 = 0.6690) was identified, whereby only compounds with exocyclic NH and
O unsaturations were excluded (Fig. 3).
The relationship between RSE and geometric parameters, including endocyclic bond distances and angles, was also examined. Thus, significant correlations were uncovered between RSEs and endocyclic C–N bond distances (Fig. S1†). For the complexed and uncomplexed azaphosphiridines with exocyclic bonds, excellent R2 values were found when represented separately (R2 = 0.9796 and 0.9999, respectively). For the remaining azaphosphiridines, an acceptable correlation of R2 = 0.6530 between the variables was found with cationic species 1e being excluded from the correlation. The notable reduction in C–N bond distances observed in compounds with exocyclic double bonds can be attributed to their partial CN double bond character, which is influenced by the participation of resonance structures (Scheme 2). Again, the importance of these structures increases with the electronegativity of the exocyclic heavy atom. Similarly, an excellent correlation was observed for the endocyclic N–P bond distances (R2 = 0.7489) when the six compounds with exocyclic unsaturations were excluded (Fig. S1†). Ultimately, the correlation with the P–C endocyclic bond distances was somewhat weaker (R2 = 0.6575) after excluding compounds containing the
NH and
O unsaturations (Fig. S1†).
On the other hand, for the case of endocyclic bond angles, a natural correspondence with the distances of the opposite bond is observed. Therefore, no meaningful correlation was identified for the C–N–P angle (Fig. S2†). However, as observed for the P–C bond distance (Fig. S1c†), the correlation strengthens when derivatives bearing exocyclic CNH, C
CH2 and C
O groups are excluded. Additionally, compound 1f was excluded from all correlations. Similarly, the P-centred bond angle correlations could be grouped into three categories, as observed for the opposite C–N bond distances (see the ESI†). Good correlations were obtained in all three cases (Fig. S2†). Finally, regarding the P–C–N angle, a similar pattern emerges as observed for the P–N opposite bond distance (Fig. 4). That is, the compounds exhibit two distinct trends, depending on whether they present (R2 = 0.881) or not (R2 = 0.634) exocyclic unsaturations (Fig. S2†).
![]() | ||
Fig. 5 ISCP of σzz (ppm) in the plane parallel to the xy (ring) plane at ±2 Å (average) for compounds (a) 1a![]() ![]() ![]() |
These results were compared with those obtained for the derivatives without exocyclic double bonds, 2a and 2b, as well as with cyclopropene 3 and the cyclopropenyl cation 4 (Fig. 6). It is evident that the cyclopropenyl cation (4) exhibits the most pronounced aromatic character among the three-membered rings displayed (Fig. 5g; the corresponding anion 4− is shown in Fig. S3†). As anticipated, cyclopropene 3 exhibits nonaromatic characteristics, whereas the unsaturated compounds 2 demonstrate the long-range effects of the substantial diatropic current generated by the lone pair on the phosphorus atom, along with a minor contribution from the CY unsaturation (the sole one present in cyclopropene 3). The latter components are only discernible in the contour maps of derivatives 1a
C, 1a
NH and 1a
O (Fig. 5).
The LUMO of oxaphosphiranes was shown to exert a pivotal influence on the approach of an incoming nucleophile,45 which ultimately resulted in the initial formation of an encounter complex.28 Similarly, in the case of azaphosphiridines, the LUMO (mostly representing the σ-hole that corresponds to σ*(P–N), as shown in Fig. S5†) directs the approach of methylamine, resulting in the formation of the encounter complex 1a·(MeNH2) (Fig. 9). The BCP for the N⋯P interaction is situated approximately at the midpoint of the bond path, as evidenced in the contour plot for the Laplacian of the electron density, ∇2ρ (Fig. 9).
![]() | ||
Fig. 9 Computed (B3LYP-D4/def2-TZVP) structure (left) for the encounter complex 1a·(MeNH2) and contour plot for the Laplacian of the electron density (B3LYP/def2-TZVPP) at the NPN plane (right). |
The most intense valence shell charge concentration (VSCC) band for the donor (D) atom, in this case N, and the shallow one for the acceptor (P) atom are located in the respective atomic basins, that is to say, at both sides of the BCP (Fig. 3). This feature is quantified by means of the adimensional parameter relative charge concentration band position,46τVSCC, defined as τVSCC = r(VSCCD)·r(VSCCP)/d2, where r(VSCC) is the (signed) position (in Å) of the VSCC band in the bond path axis relative to the BCP and ‘d’ is the internuclear distance. In the case of the encounter complex 1a·(MeNH2), the relatively high negative τVSCC value, together with the small positive ∇2ρ (Table 2), suggests that it should be classified as a van der Waals complex.46 This conclusion is further supported by the relatively long D⋯P distance and low Wiberg bond index (WBI). Additionally, it displays a moderate interaction energy, predominantly characterised by an electron transfer from the lone pair at the donor (N) atom to the endocyclic σ*(P–N) orbital, as predicted by the second-order perturbation theory (SOPT) (Table 2). Similar weak interactions (Table 2) and ∇2ρ patterns (Fig. 10) are found for the van der Waals complexes of 1a with either MeOH or PMe3. The latter complex is reminiscent of the van der Waals P⋯P complexes formed between an oxaphosphirane and PMe3 at either side of the ring P atom, which have recently been reported.45 The van der Waals complex with 1,3-dimethyl-imidazol-2-ylidene (NHCMe2) was found to be somewhat stronger than the other complexes, with the chloride complex exhibiting the greatest strength. This was evidenced by the shorter distances, higher bond orders and stronger interaction energies (Table 2).
The case of methyl amine serves to illustrate the two-fold effect of the initial van der Waals complex formation. Firstly, it increases the population of both endocyclic σ*(P–N)/σ*(P–C) orbitals (0.024/0.018 and 0.046/0.019e in 1a and 1a·(MeNH2), respectively), thereby weakening the corresponding bonds (WBIP–N/P–C = 0.891/0.960 and 0.844/0.955 for 1a and 1a·(MeNH2), respectively). Secondly, it also increases the Lagrangian kinetic energy density per electron, G(r)/ρ(r), at the ring critical point (0.977 and 1.001 au, for 1a and 1a·(MeNH2), respectively), which has been demonstrated to correlate with an increase in the RSE within a set of structurally related rings.46
Furthermore, the use of parent azaphosphiridine (1a) and its κP-W(CO)5 complex (W-1a) demonstrates that an approaching of the donor atom of the nucleophile to either the P or C ring atoms results in the promotion of P–C bond cleavage. Therefore, the energetically favourable ring opening reactions with methylamine MeNH2 of these azaphosphiridine derivatives (Scheme 3), due to their high RSEs, were investigated. The initial reaction resulted in the formation of the opening product 5, occurring via an exergonic process (ΔG° = −9.2 kcal mol−1) with a moderate activation barrier (ΔG‡ = 33.5 kcal mol−1) (Fig. 11). Conversely, the reaction of the complex initially yielded the intermediate compound W-6 (not observed in the case of uncomplexed azaphosphiridine), which exhibited a higher activation barrier (ΔG‡ = 41.4 kcal mol−1). Subsequently, the migration of hydrogen to phosphorus produced compound W-5 exergonically. For comparison purpose, the attack of methylamine on the acyclic analogues 7 and W-7 was also studied. First, the attack on 7 furnishes an ion pair intermediate (whose energy was computed as the sum of the two isolated ions) in a very endergonic process with high activation barrier (ΔG‡ = 63.7 kcal mol−1). Similarly, the attack of MeNH2 on W-7 leads to the formation of compounds 10 and W-11 in an endergonic process with a high activation barrier (ΔG‡ = 59.3 kcal mol−1). The intermediates that result from the reactions of both the cyclic and acyclic uncomplexed derivatives with methylamine are not stable, and thus do not represent energy minima. This is because the negative charge on P is not stabilised to the same extent as in the case of the complexed analogues, which are stable and could be characterised computationally as minima. The final exothermicity values (expressed in zero point-corrected energies) for the opening reactions of azaphosphiridines 1a and W-1a (ΔEZPE = −20.2 and −21.7 kcal mol−1, respectively) (Fig. S6†) roughly match the RSE of each ring at the level of calculation employed (the overall process is a lower quality isodesmic reaction, RC3).
![]() | ||
Scheme 3 Methyl amine-mediated P–C bond cleavage in parent (1a) and W(CO)5 complexed (W-1a) azaphosphiridine and model acyclic analogues 7 and W-7. |
![]() | ||
Fig. 11 Computed (DLPNO-CCSD(T)/def2-TZVPP) Gibbs free energy profile for the MeNH2-mediated P–C bond cleavage reactions depicted in Scheme 3: 1a (black); W-1a (grey); 7a (red); W-7 (blue). |
The acceptable correlation of the RSEs with the relaxed force constants of the endocyclic bonds and AIM-derived parameters such as the electron density, ρ(r), and the Lagrangian of the kinetic energy, G(r) at the ring critical points is noteworthy. It is important to highlight that the exergonicity and the relatively low barrier to the cleavage of the P–C bond by MeNH2, underlines the effect of the RSE as a driving force for the ring-opening of azaphosphiridines, especially in comparison to acyclic analogues. This could pave the way for their use in polymerisation reactions.
The data that support the findings of this study are openly available in ESI† of this manuscript.
Footnote |
† Electronic supplementary information (ESI) available: Other computational results and calculated structures. See DOI: https://doi.org/10.1039/d4dt03117b |
This journal is © The Royal Society of Chemistry 2025 |